Molecular Modeling of Surfactant Micellization ... - ACS Publications

Jan 9, 2019 - Department of Chemical and Biological Engineering, Princeton ... of the solvent-accessible surface area (SASA) of the hydrophobic domain...
5 downloads 0 Views 458KB Size
Subscriber access provided by United Arab Emirates University | Libraries Deanship

New Concepts at the Interface: Novel Viewpoints and Interpretations, Theory and Computations

Molecular Modeling of Surfactant Micellization Using Solvent Accessible Surface Area Hsieh Chen, and Athanassios Z. Panagiotopoulos Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.8b03440 • Publication Date (Web): 09 Jan 2019 Downloaded from http://pubs.acs.org on January 14, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Molecular Modeling of Surfactant Micellization Using Solvent Accessible Surface Area Hsieh Chen1,* and Athanassios Z. Panagiotopoulos2 1Aramco

Services Company: Aramco Research Center – Boston, 400 Technology Square, Cambridge,

MA 02139 2Department

of Chemical and Biological Engineering, Princeton University, Princeton, NJ 08544

*[email protected]

1 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 24

ABSTRACT We report a new implicit-solvent simulation model for studying the self-assembly of surfactants, where the hydrophobic interactions were captured by calculating the relative changes of the solvent accessible surface area (SASA) of the hydrophobic domains. Using histogram-reweighting grand canonical Monte Carlo simulations, we demonstrate that this approach allows us to match both the experimental critical micelle concentrations (cmc) and micellar aggregation numbers simultaneously with a single phenomenological surface tension γSASA for the poly(oxyethylene) monoalkyl ether (CmEn) surfactants in aqueous solutions. Excellent transferability is observed: the same model can accurately predict the experimental cmc and aggregation numbers for the CmEn surfactants with alkyl lengths m between 6 and 12, and poly(oxyethylene) lengths n between 1 and 9. The SASA-based implicit-solvent model put forward in this work is general and may be applied to study more complex amphiphilic systems such as surfactants with branched alkyl chains, or surfactant-hydrocarbon mixtures.

2 ACS Paragon Plus Environment

Page 3 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

INTRODUCTION Surfactants are present in a large number of industrial applications, including detergency, catalysis, pharmaceutical, food, cosmetic, materials synthesis, nanotechnology, and tertiary oil recovery.17

In nature, lipids are the most important components of biological cell membranes and are involved in

multiple biochemistry processes.8, 9 The key features of the surfactant molecules are their ability to selfassemble into a variety of nanostructures, such as spherical and cylindrical micelles, hexagonal phases, cubic phases, and lamellae. Moreover, their morphologies are usually controllable by the different surfactant molecular architectures, concentrations, temperatures, solvent types, co-surfactants, and the presence of salts or other solutes.10-12 Because of its importance as well as its complexity, micellization in surfactant solutions has been extensively studied by experiments,13-17 theories,18-23 and computer simulations.24-34 Molecule-based simulations are particularly useful in providing direct links between the molecular scale interactions and the diverse self-assembly behaviors. Explicit-solvent molecular dynamics simulations have been widely used to study the structural properties, water penetration, and counterion bindings of (usually preassembled) micelles.35-42 However, due to computational limitations, it has been difficult to simulate micellization processes using explicit-solvent models.43-46 Typical micellization time scales (>1 μs)47 can only be reached through extensive simulations. Even worse, micellization usually occurs at low concentrations ( 300 before the termination of the specific simulation run, indicating phase separation instead of micellization of the surfactants.

CONCLUSIONS In this work, we have demonstrated that calculating the surface tensions associated with the hydrophobic solvent accessible surface area (SASA) is a suitable approach for capturing the hydrophobic interactions in surfactant micellization. Specifically, we applied the model to study the nonionic poly(oxyethylene) ether surfactant systems CmEn, with m denoting the hydrophobic alkyl chain lengths and n denoting the hydrophilic poly(oxyethylene) chain lengths. We first adjusted the surface tension 9 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 24

γSASA to match the critical micelle concentration (cmc) and micellar aggregation number (M) of the C8E6 surfactants. Encouragingly, both cmc and M matched the experimental values with the same γSASA/kB = 32.5 K/Å2 (γSASA = 65 cal/mol/Å2), which is a realistic value in between the reported microscopic (28-47 cal/mol/Å2) and macroscopic (72-75 cal/mol/Å2) hydrocarbon-water interface tensions. We then used the optimized γSASA to predict the cmc and M for the different CmEn surfactants with varied m = 6 to 12 and n = 1 to 9, and found very good agreements with the respective experimental values, indicating excellent transferability of the present method. Finally, we investigated the snapshots of the isolated micellar aggregates from simulations and found spherical micelles for the C6E6, C8E6, and C8E9 systems; ellipsoidal micelles for the C8E3, C10E6, and C12E6 systems; and the phase separated sheet aggregate for the C8E1 system. The observations of the different micellar shapes were consistent with prior theories and simulations, which highlighted the validity and generality of the new implicit-solvent surfactant model put forward in this work. Even though in this work we have mainly focused on nonionic surfactants, the SASA-based model may be readily generalized to study ionic surfactants as well, since the SASA-based hydrophobic tail bead interaction is agnostic to the surfactant head groups. We then would only need to acquire the parameters for new head groups without modifying SASA parameters. Previously, we have developed an implicit-solvent molecular model for electrolytes based on the concept of concentration dependent dielectric permittivity, which was demonstrated to precisely capture the thermodynamic properties of concentrated electrolytes and their mixtures.66 We believe the same method may be applied to describe the charged head groups of the ionic surfactants; further, we may combine the surfactant and electrolyte models to study the ion-specific effects in micellization. A final point relates to the temperature-dependent properties. In this work, we have obtained the optimized γSASA at T = 298.15 K. Nevertheless, it has been known that the water-hydrocarbon interface tension is a non-linear function of temperature, which is governed by the collective hydration behavior.109, 110

The temperature-dependent properties may be captured by first obtaining γSASA(T) either by adjusting it 10 ACS Paragon Plus Environment

Page 11 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

at different temperatures, or by applying other analytical corrections. With optimal γSASA(T), we expect our model can also be used to capture the temperature driven phenomena such as the cloud points for the CmEn surfactants. Further studies in this area are needed to validate the aforementioned approaches.

ACKNOWLEDGEMENTS We thank Aramco Research Center-Boston, Reservoir Engineering Technology team members for valuable discussions. We acknowledge ASC, IT Services for managing computational resources.

REFERENCES (1) Karsa, D. R., Industrial Applications of Surfactants IV. The Royal Society of Chemistry: Cambridge, UK, 1999. (2) Kumar, D.; Seth, K.; Kommi, D. N.; Bhagat, S.; Chakraborti, A. K., Surfactant Micelles as Microreactors for the Synthesis of Quinoxalines in Water: Scope and Limitations of Surfactant Catalysis. RSC Adv. 2013, 3, 15157-15168. (3) Van Lehn, R.; Sing, C.; Chen, H.; Alexander-Katz, A., Multidimensional Targeting: Using Physical and Chemical Forces in Unison. Curr. Pharm. Biotechnol. 2010, 11, 320-332. (4) Huo, Q.; Margolese, D. I.; Stucky, G. D., Surfactant Control of Phases in the Synthesis of Mesoporous Silica-based Materials. Chem. Mater. 1996, 8, 1147-1160. (5) Mann, S.; Ozin, G. A., Synthesis of Inorganic Materials with Complex Form. Nature 1996, 382, 313. (6) Hirasaki, G.; Miller, C. A.; Puerto, M., Recent Advances in Surfactant EOR. SPE J. 2011, 16, 889-907. (7) Lake, L. W.; Johns, R. T.; Rossen, W. R.; Pope, G. A., Fundamentals of Enhanced Oil Recovery. Society of Petroleum Engineers: Richardson, TX, 2014. (8) Nagle, J. F.; Tristram-Nagle, S., Structure of Lipid Bilayers. Biochim. Biophys. Acta 2000, 1469, 159-195. (9) Gurr, M. I.; Harwood, J. L.; Frayn, K. N., Lipid Biochemistry. Springer: Boston, MA, 2002. (10) Tanford, C., The Hydrophobic Effect: Formation of Micelles and Biological Membranes 2d Edition. John Wiley & Sons: Somerset, NJ, 1980. (11) Israelachvili, J. N., Intermolecular and Surface Forces. Elsevier: Waltham, MA, 2011. (12) Rosen, M. J.; Kunjappu, J. T., Surfactants and Interfacial Phenomena. John Wiley & Sons: Hoboken, NJ, 2012. (13) Zana, R., Aqueous Surfactant-Alcohol Systems: A Review. Adv. Colloid Interface Sci. 1995, 57, 1-64. (14) Griffiths, P.; Stilbs, P.; Paulsen, K.; Howe, A.; Pitt, A., FT-PGSE NMR Study of Mixed Micellization of an Anionic and a Sugar-based Nonionic Surfactant. J. Phys. Chem. B 1997, 101, 915918. 11 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 24

(15) Carpena, P.; Aguiar, J.; Bernaola-Galván, P.; Carnero Ruiz, C., Problems Associated with the Treatment of Conductivity−Concentration Data in Surfactant Solutions: Simulations and Experiments. Langmuir 2002, 18, 6054-6058. (16) Bouchemal, K.; Agnely, F.; Koffi, A.; Djabourov, M.; Ponchel, G., What Can Isothermal Titration Microcalorimetry Experiments Tell Us about the Self‐organization of Surfactants into Micelles? J. Mol. Recognit. 2010, 23, 335-342. (17) Dong, R.; Hao, J., Complex Fluids of Poly(oxyethylene) Monoalkyl Ether Nonionic Surfactants. Chem. Rev. 2010, 110, 4978-5022. (18) Kahlweit, M.; Teubner, M., On the Kinetics of Micellization in Aqueous Solutions. Adv. Colloid Interface Sci. 1980, 13, 1-64. (19) Puvvada, S.; Blankschtein, D., Molecular‐Thermodynamic Approach to Predict Micellization, Phase Behavior and Phase Separation of Micellar Solutions. I. Application to Nonionic Surfactants. J. Chem. Phys. 1990, 92, 3710-3724. (20) Hinze, W. L.; Pramauro, E., A Critical Review of Surfactant-mediated Phase Separations (Cloudpoint Extractions): Theory and Applications. Crit. Rev. Anal. Chem. 1993, 24, 133-177. (21) Zana, R., Critical Micellization Concentration of Surfactants in Aqueous Solution and Free Energy of Micellization. Langmuir 1996, 12, 1208-1211. (22) Boström, M.; Williams, D. R.; Ninham, B. W., Ion Specificity of Micelles Explained by Ionic Dispersion Forces. Langmuir 2002, 18, 6010-6014. (23) Srinivasan, V.; Blankschtein, D., Effect of Counterion Binding on Micellar Solution Behavior: 1. Molecular− Thermodynamic Theory of Micellization of Ionic Surfactants. Langmuir 2003, 19, 99329945. (24) Palmer, B. J.; Liu, J., Simulations of Micelle Self-Assembly in Surfactant Solutions. Langmuir 1996, 12, 746-753. (25) Marrink, S.; Tieleman, D.; Mark, A., Molecular Dynamics Simulation of the Kinetics of Spontaneous Micelle Formation. J. Phys. Chem. B 2000, 104, 12165-12173. (26) Shelley, J. C.; Shelley, M. Y., Computer Simulation of Surfactant Solutions. Curr. Opin. Colloid Interface Sci. 2000, 5, 101-110. (27) Bandyopadhyay, S.; Tarek, M.; Lynch, M. L.; Klein, M. L., Molecular Dynamics Study of the Poly (oxyethylene) Surfactant C12E2 and Water. Langmuir 2000, 16, 942-946. (28) Sangwai, A. V.; Sureshkumar, R., Coarse-grained Molecular Dynamics Simulations of the Sphere to Rod Transition in Surfactant Micelles. Langmuir 2011, 27, 6628-6638. (29) Cheong, D. W.; Panagiotopoulos, A. Z., Monte Carlo Simulations of Micellization in Model Ionic Surfactants: Application to Sodium Dodecyl Sulfate. Langmuir 2006, 22, 4076-4083. (30) Sanders, S. A.; Panagiotopoulos, A. Z., Micellization Behavior of Coarse Grained Srfactant Mdels. J. Chem. Phys. 2010, 132, 114902. (31) Sammalkorpi, M.; Sanders, S.; Panagiotopoulos, A.; Karttunen, M.; Haataja, M., Simulations of Micellization of Sodium Hexyl Sulfate. J. Phys. Chem. B 2011, 115, 1403-1410. (32) Sanders, S. A.; Sammalkorpi, M.; Panagiotopoulos, A. Z., Atomistic Simulations of Micellization of Sodium Hexyl, Heptyl, Octyl, and Nonyl Sulfates. J. Phys. Chem. B 2012, 116, 2430-2437. (33) Jusufi, A.; Panagiotopoulos, A. Z., Explicit-and Implicit-Solvent Simulations of Micellization in Surfactant Solutions. Langmuir 2014, 31, 3283-3292. (34) Jiao, S.; Santos, A. P.; Panagiotopoulos, A. Z., Differences in Free Surfactant Concentration and Aggregation Properties for Amphiphiles with the Same Critical Micelle Concentration. Fluid Phase Equilib. 2018, 470, 126-133. (35) MacKerell Jr, A. D., Molecular Dynamics Simulation Analysis of a Sodium Dodecyl Sulfate Micelle in Aqueous Solution: Decreased Fluidity of the Micelle Hydrocarbon Interior. J. Phys. Chem. 1995, 99, 1846-1855. (36) Sterpone, F.; Briganti, G.; Pierleoni, C., Molecular Dynamics Study of Spherical Aggregates of Chain Molecules at Different Degrees of Hydrophilicity in Water Solution. Langmuir 2001, 17, 51035110. 12 ACS Paragon Plus Environment

Page 13 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(37) Bruce, C. D.; Berkowitz, M. L.; Perera, L.; Forbes, M. D., Molecular Dynamics Simulation of Sodium Dodecyl Sulfate Micelle in Water: Micellar Structural Characteristics and Counterion Distribution. J. Phys. Chem. B 2002, 106, 3788-3793. (38) Bruce, C. D.; Senapati, S.; Berkowitz, M. L.; Perera, L.; Forbes, M. D., Molecular Dynamics Simulations of Sodium Dodecyl Sulfate Micelle in Water: the Behavior of Water. J. Phys. Chem. B 2002, 106, 10902-10907. (39) Garde, S.; Yang, L.; Dordick, J. S.; Paulaitis, M. E., Molecular Dynamics Simulation of C8E5 Micelle in Explicit Water: Structure and Hydrophobic Solvation Thermodynamics. Mol. Phys. 2002, 100, 2299-2306. (40) Rakitin, A. R.; Pack, G. R., Molecular Dynamics Simulations of Ionic Interactions with Dodecyl Sulfate Micelles. J. Phys. Chem. B 2004, 108, 2712-2716. (41) Yoshii, N.; Iwahashi, K.; Okazaki, S., A Molecular Dynamics Study of Free Energy of Micelle Formation for Sodium Dodecyl Sulfate in Water and its Size Distribution. J. Chem. Phys. 2006, 124, 184901. (42) Sterpone, F.; Briganti, G.; Pierleoni, C., Sphere versus Cylinder: The Effect of Packing on the Structure of Nonionic C12E6 Micelles. Langmuir 2009, 25, 8960-8967. (43) Mohan, G.; Kopelevich, D. I., A Multiscale Model for Kinetics of Formation and Disintegration of Spherical Micelles. J. Chem. Phys. 2008, 128, 044905. (44) Vishnyakov, A.; Lee, M.-T.; Neimark, A. V., Prediction of the Critical Micelle Concentration of Nonionic Surfactants by Dissipative Particle Dynamics Simulations. J. Phys. Chem. Lett. 2013, 4, 797802. (45) Jakobtorweihen, S.; Yordanova, D.; Smirnova, I., Predicting Critical Micelle Concentrations with Molecular Dynamics Simulations and COSMOmic. Chem. Ing. Tech. 2017, 89, 1288-1296. (46) Ruiz-Morales, Y.; Romero-Martínez, A., Coarse-Grain Molecular Dynamics Simulations To Investigate the Bulk Viscosity and Critical Micelle Concentration of the Ionic Surfactant Sodium Dodecyl Sulfate (SDS) in Aqueous Solution. J. Phys. Chem. B 2018, 122, 3931-3943. (47) Alexandridis, P.; Holzwarth, J. F.; Hatton, T. A., Interfacial Dynamics of Water-in-Oil Microemulsion Droplets: Determination of the Bending Modulus Using Iodine Laser Temperature Jump. Langmuir 1993, 9, 2045-2052. (48) Huibers, P. D.; Lobanov, V. S.; Katritzky, A. R.; Shah, D. O.; Karelson, M., Prediction of Critical Micelle Concentration Using a Quantitative Structure− Property Relationship Approach. 1. Nonionic Surfactants. Langmuir 1996, 12, 1462-1470. (49) Huibers, P. D.; Lobanov, V. S.; Katritzky, A.; Shah, D.; Karelson, M., Prediction of Critical Micelle Concentration Using a Quantitative Structure–Property Relationship Approach. 2. Anionic Surfactants. J. Colloid Interface Sci. 1997, 187, 113-120. (50) le Maire, M.; Champeil, P.; Møller, J. V., Interaction of Membrane Proteins and Lipids with Solubilizing Detergents. Biochim. Biophys. Acta 2000, 1508, 86-111. (51) Shinoda, W.; DeVane, R.; Klein, M. L., Multi-Property Fitting and Parameterization of a Coarse Grained Model for Aqueous Surfactants. Mol. Simul. 2007, 33, 27-36. (52) Shinoda, W.; DeVane, R.; Klein, M. L., Coarse-Grained Molecular Modeling of Non-Ionic Surfactant Self-Assembly. Soft Matter 2008, 4, 2454-2462. (53) Drouffe, J.; Maggs, A.; Leibler, S., Computer Simulations of Self-Assembled Membranes. Science 1991, 254, 1353-1356. (54) Shinto, H.; Morisada, S.; Miyahara, M.; Higashitani, K., Langevin Dynamics Simulations of Cationic Surfactants in Aqueous Solutions Using Potentials of Mean Force. Langmuir 2004, 20, 20172025. (55) Brannigan, G.; Tamboli, A. C.; Brown, F. L., The Role of Molecular Shape in Bilayer Elasticity and Phase Behavior. J. Chem. Phys. 2004, 121, 3259-3271. (56) Lazaridis, T.; Mallik, B.; Chen, Y., Implicit Solvent Simulations of DPC Micelle Formation. J. Phys. Chem. B 2005, 109, 15098-15106. 13 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 24

(57) Morisada, S.; Shinto, H.; Higashitani, K., Revised Implicit Solvent Model for the Simulation of Surfactants in Aqueous Solutions. J. Phys. Chem. B 2005, 109, 11762-11769. (58) Morisada, S.; Shinto, H.; Higashitani, K., Revised Implicit Solvent Model for the Simulation of Surfactants in Aqueous Solutions. 2. Modeling of Charged Headgroups at Oil− Water Interface. J. Chem. Theory Comput. 2007, 3, 1163-1171. (59) Allen, E. C.; Rutledge, G. C., A Novel Algorithm for Creating Coarse-Grained, Density Dependent Implicit Solvent Models. J. Chem. Phys. 2008, 128, 154115. (60) Izvekov, S.; Voth, G. A., Solvent-Free Lipid Bilayer Model Using Multiscale Coarse-Graining. J. Phys. Chem. B 2009, 113, 4443-4455. (61) Morisada, S.; Shinto, H., Implicit Solvent Model Simulations of Surfactant Self-Assembly in Aqueous Solutions. J. Phys. Chem. B 2010, 114, 6337-6343. (62) Wang, Z.-J.; Deserno, M., Systematic Implicit Solvent Coarse-Graining of Bilayer Membranes: Lipid and Phase Transferability of the Force Field. New J. Phys. 2010, 12, 095004. (63) Kim, Y.; Alexander-Katz, A., Phase Behavior of Symmetric Disk-Coil Macromolecules with Stacking Interactions. J. Chem. Phys. 2011, 135, 024902. (64) Kim, Y.; Ha, E.; Alexander-Katz, A., Phase Behavior of Disk–Coil Macromolecules. Macromolecules 2011, 44, 7016-7025. (65) Wang, S.; Larson, R. G., A Coarse-Grained Implicit Solvent Model for Poly (ethylene oxide), CnEm Surfactants, and Hydrophobically End-Capped Poly (ethylene oxide) and Its Application to Micelle Self-Assembly and Phase Behavior. Macromolecules 2015, 48, 7709-7718. (66) Chen, H.; Panagiotopoulos, A. Z., Communication: Modeling Electrolyte Mixtures with Concentration Dependent Dielectric Permittivity. J. Chem. Phys. 2018, 148, 041102. (67) Jusufi, A.; Hynninen, A.-P.; Panagiotopoulos, A. Z., Implicit Solvent Models for Micellization of Ionic Surfactants. J. Phys. Chem. B 2008, 112, 13783-13792. (68) Jusufi, A.; Hynninen, A.-P.; Haataja, M.; Panagiotopoulos, A. Z., Electrostatic Screening and Charge Correlation Effects in Micellization of Ionic Surfactants. J. Phys. Chem. B 2009, 113, 6314-6320. (69) Jusufi, A.; Sanders, S.; Klein, M. L.; Panagiotopoulos, A. Z., Implicit-Solvent Models for Micellization: Nonionic Surfactants and Temperature-Dependent Properties. J. Phys. Chem. B 2011, 115, 990-1001. (70) Reed, R. L.; Healy, R. N., Some Physicochemical Aspects of Microemulsion Flooding: A Review. In Improved Oil Recovery by Surfactant and polymer flooding, Academic Press: New York, NY, 1977; pp 383-437. (71) Kumar, P.; Mittal, K. L., Handbook of Microemulsion Science and Technology. Marcel Dekker: New York, NY, 1999. (72) Muzaffar, F.; Singh, U.; Chauhan, L., Review on Microemulsion as Futuristic Drug Delivery. Int. J. Pharm. Pharm. Sci. 2013, 5, 39-53. (73) Chandler, D., Interfaces and the Driving Force of Hydrophobic Assembly. Nature 2005, 437, 640. (74) Lee, B.; Richards, F. M., The Interpretation of Protein Structures: Estimation of Static Accessibility. J. Mol. Biol. 1971, 55, 379-IN4. (75) Shrake, A.; Rupley, J., Environment and Exposure to Solvent of Protein Atoms. Lysozyme and Insulin. J. Mol. Biol. 1973, 79, 351-371. (76) Nicholls, A.; Sharp, K. A.; Honig, B., Protein Folding and Association: Insights from the Interfacial and Thermodynamic Properties of Hydrocarbons. Proteins 1991, 11, 281-296. (77) Marsh, J. A.; Teichmann, S. A., Relative Solvent Accessible Surface Area Predicts Protein Conformational Changes upon Binding. Structure 2011, 19, 859-867. (78) White, S. H.; Wimley, W. C., Membrane Protein Folding and Stability: Physical Principles. Annu. Rev. Biophys. Biomol. Struct. 1999, 28, 319-365. (79) Choe, S.; Hecht, K. A.; Grabe, M., A Continuum Method for Determining Membrane Protein Insertion Energies and the Problem of Charged Residues. J. Gen. Physiol. 2008, 131, 563-573.

14 ACS Paragon Plus Environment

Page 15 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(80) Mondal, S.; Khelashvili, G.; Shan, J.; Andersen, O. S.; Weinstein, H., Quantitative Modeling of Membrane Deformations by Multihelical Membrane Proteins: Application to G-Protein Coupled Receptors. Biophys. J. 2011, 101, 2092-2101. (81) Öjemalm, K.; Higuchi, T.; Jiang, Y.; Langel, Ü.; Nilsson, I.; White, S. H.; Suga, H.; von Heijne, G., Apolar Surface Area Determines the Efficiency of Translocon-Mediated Membrane-Protein Integration into the Endoplasmic Reticulum. Proc. Natl. Acad. Sci. USA 2011, 108, E359-E364. (82) Van Lehn, R. C.; Atukorale, P. U.; Carney, R. P.; Yang, Y.-S.; Stellacci, F.; Irvine, D. J.; Alexander-Katz, A., Effect of Particle Diameter and Surface Composition on the Spontaneous Fusion of Monolayer-Protected Gold Nanoparticles with Lipid Bilayers. Nano Lett. 2013, 13, 4060-4067. (83) Van Lehn, R. C.; Alexander-Katz, A., Ligand-Mediated Short-Range Attraction Drives Aggregation of Charged Monolayer-Protected Gold Nanoparticles. Langmuir 2013, 29, 8788-8798. (84) Van Lehn, R. C.; Alexander-Katz, A., Free Energy Change for Insertion of Charged, MonolayerProtected Nanoparticles into Lipid Bilayers. Soft Matter 2014, 10, 648-658. (85) Van Lehn, R. C.; Alexander-Katz, A., Fusion of Ligand-Coated Nanoparticles with Lipid Bilayers: Effect of Ligand Flexibility. J. Phys. Chem. A 2014, 118, 5848-5856. (86) Jorgensen, W. L.; Madura, J. D.; Swenson, C. J., Optimized Intermolecular Potential Functions for Liquid Hydrocarbons. J. Am. Chem. Soc. 1984, 106, 6638-6646. (87) Martin, M. G.; Siepmann, J. I., Transferable Potentials for Phase Equilibria. 1. United-Atom description of n-Alkanes. J. Phys. Chem. B 1998, 102, 2569-2577. (88) Weeks, J. D.; Chandler, D.; Andersen, H. C., Role of Repulsive Forces in Determining the Equilibrium Structure of Simple Liquids. J. Chem. Phys. 1971, 54, 5237-5247. (89) Sharp, K. A.; Nicholls, A.; Fine, R. F.; Honig, B., Reconciling the Magnitude of the Microscopic and Macroscopic Hydrophobic Effects. Science 1991, 252, 106-109. (90) Sitkoff, D.; Ben-Tal, N.; Honig, B., Calculation of Alkane to Water Solvation Free Energies Using Continuum Solvent Models. J. Phys. Chem. 1996, 100, 2744-2752. (91) Huang, D. M.; Geissler, P. L.; Chandler, D., Scaling of Hydrophobic Solvation Free Energies. J. Phys. Chem. B 2001, 105, 6704-6709. (92) Huang, D. M.; Chandler, D., The Hydrophobic Effect and the Influence of Solute−Solvent Attractions. J. Phys. Chem. B 2002, 106, 2047-2053. (93) Panagiotopoulos, A. Z.; Wong, V.; Floriano, M. A., Phase Equilibria of Lattice Polymers from Histogram Reweighting Monte Carlo Simulations. Macromolecules 1998, 31, 912-918. (94) Floriano, M. A.; Caponetti, E.; Panagiotopoulos, A. Z., Micellization in Model Surfactant Systems. Langmuir 1999, 15, 3143-3151. (95) Panagiotopoulos, A. Z.; Floriano, M. A.; Kumar, S. K., Micellization and Phase Separation of Diblock and Triblock Model Surfactants. Langmuir 2002, 18, 2940-2948. (96) Panagiotopoulos, A. Z., Monte Carlo Methods for Phase Equilibria of Fluids. J. Phys.: Condens. Matter 2000, 12, R25. (97) Ferrenberg, A. M.; Swendsen, R. H., New Monte Carlo Technique for Studying Phase Transitions. Phys. Rev. Lett. 1988, 61, 2635. (98) Ferrenberg, A. M.; Swendsen, R. H., Optimized Monte Carlo Data Analysis. Phys. Rev. Lett. 1989, 63, 1195. (99) Shah, J. K.; Maginn, E. J., A General and Efficient Monte Carlo Method for Sampling Intramolecular Degrees of Freedom of Branched and Cyclic Molecules. J. Chem. Phys. 2011, 135, 134121. (100) Swinbank, R.; Purser, R. J., Fibonacci Grids: A Novel Approach to Global Modelling. Q. J. R. Meteorol. Soc. 2006, 132, 1769-1793. (101) Eisenhaber, F.; Lijnzaad, P.; Argos, P.; Sander, C.; Scharf, M., The Double Cubic Lattice Method: Efficient Approaches to Numerical Integration of Surface Area and Volume and to Dot Surface Contouring of Molecular Assemblies. J. Comput. Chem. 1995, 16, 273-284. (102) Mitternacht, S., FreeSASA: An Open Source C Library for Solvent Accessible Surface Area Calculations. arXiv:1601.06764 2016. 15 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 24

(103) Santos, A. P.; Panagiotopoulos, A. Z., Determination of the Critical Micelle Concentration in Simulations of Surfactant Systems. J. Chem. Phys. 2016, 144, 044709. (104) Corkill, J.; Goodman, J.; Tate, J., Calorimetric Determination of the Heats of Micelle Formation of Some Non-ionic Detergents. Trans. Faraday Soc. 1964, 60, 996-1002. (105) Shinoda, K.; Yamanaka, T.; Kinoshita, K., Surface Chemical Properties in Aqueous Solutions of Non-ionic Surfactants Octyl Glycol Ether, α-Octyl Glyceryl Ether and Octyl Glucoside. J. Phys. Chem. 1959, 63, 648-650. (106) Corkill, J.; Goodman, J.; Harrold, S., Thermodynamics of Micellization of Non-Ionic Detergents. Trans. Faraday Soc. 1964, 60, 202-207. (107) Tanford, C., Micelle Shape and Size. J. Phys. Chem. 1972, 76, 3020-3024. (108) Robson, R. J.; Dennis, E. A., The Size, Shape, and Hydration of Nonionic Surfactant Micelles. Triton X-100. J. Phys. Chem. 1977, 81, 1075-1078. (109) Chen, H.; Cox, J. R.; Ow, H.; Shi, R.; Panagiotopoulos, A. Z., Hydration Repulsion between Carbohydrate Surfaces Mediated by Temperature and Specific Ions. Sci. Rep. 2016, 6, 28553. (110) Jiménez-Ángeles, F.; Firoozabadi, A., Hydrophobic Hydration and the Effect of NaCl Salt in the Adsorption of Hydrocarbons and Surfactants on Clathrate Hydrates. ACS Cent. Sci. 2018, 4, 820-831.

16 ACS Paragon Plus Environment

Page 17 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 1. Schematic illustration that the hydrophobic interactions were captured from the surface tensions associated with the hydrophobic solvent accessible surface area (SASA).

17 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 24

Figure. 2. (A) Osmotic pressure (Π = P) versus concentration (c = N/V) curves, and (B) aggregate size distributions (with M denoting micellar aggregation numbers) for the C8E6 surfactants with different γSASA. The black line in (A) corresponds to ideal solutions.

18 ACS Paragon Plus Environment

Page 19 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure. 3. (A) Critical micelle concentrations (cmc) and (B) micellar aggregation numbers (M) as a function of γSASA for the C8E6 surfactants from implicit-solvent GCMC simulations (black squares). Experimental values from Ref. 50 are shown as red dashed lines.

19 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

Figure 4. (A) Critical micelle concentrations (cmc) and (B) micellar aggregation numbers (M) for the CmE6 surfactants from implicit-solvent GCMC simulations (black squares) and from experiments (red triangles).50, 104 The red dashed line in (A) is an exponential fit for the experimental results. Error bars are only shown if larger than the symbols.

20 ACS Paragon Plus Environment

Page 21 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 5. (A) Critical micelle concentrations (cmc) and (B) micellar aggregation numbers (M) for the C8En surfactants from implicit-solvent GCMC simulations (black squares) and from experiments (red triangles).50, 105, 106 The red dashed line in (A) is a linear fit for the experimental results. Error bars are only shown if larger than the symbols.

21 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 24

Figure 6. Snapshots of the isolated micelles from the different CmEn surfactant systems studied in this work. The hydrophobic beads are shown in cyan, and the hydrophilic beads are shown in orange.

22 ACS Paragon Plus Environment

Page 23 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 7. Snapshots of the phase separated C8E1 surfactant aggregate. The bead colors are the same as in Fig. 6.

23 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 24

For Table of Contents Use Only.

24 ACS Paragon Plus Environment